You are on page 1of 18

Copyright 2000 by the Genetics Society of America

Empirical Bayes Procedure for Estimating Genetic Distance Between Populations


and Effective Population Size
Shuichi Kitada,* Takeshi Hayashi and Hirohisa Kishino
*Department of Aquatic Biosciences, Tokyo University of Fisheries, Minato, Tokyo 108-8477, Japan and Graduate School of Agricultural
and Life Sciences, University of Tokyo, Bunkyo, Tokyo 113-8657, Japan
Manuscript received February 17, 2000
Accepted for publication July 26, 2000
ABSTRACT
We developed an empirical Bayes procedure to estimate genetic distances between populations using
allele frequencies. This procedure makes it possible to describe the skewness of the genetic distance while
taking full account of the uncertainty of the sample allele frequencies. Dirichlet priors of the allele
frequencies are specified, and the posterior distributions of the various composite parameters are obtained
by Monte Carlo simulation. To avoid overdependence on subjective priors, we adopt a hierarchical model
and estimate hyperparameters by maximizing the joint marginal-likelihood function. Taking advantage
of the empirical Bayesian procedure, we extend the method to estimate the effective population size using
temporal changes in allele frequencies. The method is applied to data sets on red sea bream, herring,
northern pike, and ayu broodstock. It is shown that overdispersion overestimates the genetic distance and
underestimates the effective population size, if it is not taken into account during the analysis. The joint
marginal-likelihood function also estimates the rate of gene flow into island populations.

S a stock management tool to counteract decreased


or depleted fishery resources, stock enhancement
programs have been undertaken in many countries for
salmonid (Hilborn and Winton 1993; Ritter 1997;
Kaeriyama 1999; Knapp 1999) and for other marine
species (Bartley 1999; Kitada 1999). Concerns about
the genetic effects of hatchery releases on wild populations have increased and aroused discussion (Walters
1988; Waples 1991; Hilborn 1992; Utter 1998;
Waples 1999). Campton (1995) reviewed the genetic
effects of hatchery releases on natural stocks of salmon
and brown trout and concluded that the empirical data
supporting those arguments are absent or largely circumstantial. This is a complex topic that needs further
research (Waples 1999). A 10-point approach for a
responsible stock enhancement program has been proposed, which includes the need to use genetic resource
management to avoid deleterious genetic effects (Blankenship and Leber 1995). Using wild individuals as
broodstock may possibly reduce genetic risks (Bartley
et al. 1995; Harada et al. 1998).
The genetic identity between produced progenies and
the wild stock will be required before one can release the
progenies. To examine the genetic identity, statistically
significant differences are required. The homogeneity
2 test of allele frequencies is commonly used for testing
genetic differences and the Roff test (Roff and Bentzen
1989) is used when minor alleles exist. If the null hypoth-

Corresponding author: Shuichi Kitada, Tokyo University of Fisheries,


4-5-7 Konan, Minato, Tokyo 108-8477, Japan.
E-mail: kitada@tokyo-u-fish.ac.jp
Genetics 156: 20632079 (December 2000)

esis is not rejected, the statistical power is required to


be reported from a conservation viewpoint (Peterman
1990; Dizon et al. 1995). However, when the genetic
difference is small, the corresponding statistical power
may also be small with small sample sizes, making it
difficult to conclude that there is no genetic difference.
The statistical power is the probability of detecting the
alternative hypothesis when it is correct. A considerably
large sample size is required if one wants to obtain
satisfactory large statistical power to reject the null hypothesis and detect small genetic differences. The hypothesis testing framework implies rejecting the null
hypothesis, so it does not work well for detecting the
genetic identity, and calculating the power is meaningless. Problems of the null hypothesis testing framework
are discussed in Cohen (1994) and Hagen (1997).
An effective method of determining genetic identity
is to examine the genetic distances between populations. If an estimated confidence interval of the genetic
distance between two populations includes 0, we could
conclude that the populations are genetically identical
or not statistically significantly different. There are several measures for the genetic distance (Nei 1987). However, the sample distributions of these genetic distances
are unknown. It is then inappropriate to estimate the
confidence intervals of the genetic distance using asymptotic variances of the estimator.
In this article, we develop a Bayesian estimating procedure to measure genetic distances between populations
from allele frequencies. We can directly evaluate the
probability distribution of the genetic distance from its
posterior distribution. The general method developed

2064

S. Kitada, T. Hayashi and H. Kishino

here is extended to estimate the effective population


size termed Ne in a population based on temporal
changes in allele frequencies. The joint marginal-likelihood function derived here coincides with the likelihood function to estimate the rate of gene flow into
island populations using the sample allele frequency
from a number of islands (Rannala and Hartigan
1996).

METHODS

Let the frequencies of k alleles of two populations to


be compared be p11, , p1k and p21, , p2k. Sanghvi
(1953) proposed that the genetic distance between two
populations be determined by
D

2(p1i p2i)2
.
i1 p1i p2i
k

(1)

We use this distance as a natural measure of the genetic


distance between populations, which takes values be /(n1. n2.)
tween 0 and 4. It is known that 2n1.n2.D
follows a 2 distribution with degree of freedom k 1
when p1i p2i pi for i 1, , k (Nei 1987), where
n1. and n2. are sample sizes (individuals) of the two
is the estimator obtained by substipopulations, and D
tuting sample frequencies in Equation 1. However, the
is unknown when p1i p2i for i
distribution of D
1, , k. It is then inappropriate to evaluate the confi,
dence interval of D using an asymptotic variance of D
although it can be derived. Here we directly evaluate
the posterior probability density of the genetic distance
measure using a Bayesian framework.
Prior and posterior distribution of D: It is not easy to
describe a reasonable prior distribution of D, especially
when we compare more than two populations. Alternatively, we set a prior for allele frequencies. Let the allele
frequency of a population be p (p1, , pk) and the
sample count be n (n1, , nk), where Rpi 1 and
Rni 2n (n individuals). When the sample is collected
by a simple random sampling procedure with replacement, n follows a multinomial distribution. A distribution is known as a conjugate prior of the binomial parameter p. A Dirichlet distribution is a conjugate prior
of multinomial proportions, which is an extension of a
distribution (Johnson and Kotz 1969; Lee 1989):
(p|)

(Rki1i) k i1
pi .
ki1(i) i1

(2)

Here, (1, , k) are regarded as hyperparameters specifying the prior distribution. We use this distribution as a prior for allele frequencies.
The posterior distribution is obtained by multiplying
the likelihood function, which is multinomial distribution in this case, by the prior. The posterior distribution
of p is then given by

P(p|n)

(Rki1ni)! (Rki1i) k ini1


pi
ki1ni! ki1(i)i1
(Rki1(i ni)) k ini1
,
pi
ki1(i ni) i1

which is again a Dirichlet distribution with parameters


modified by the data ni i (Lange 1995; Weir 1996).
Given (1, , k), we can obtain a posterior
distribution of p by generating Dirichlet random numbers with parameter n using Monte Carlo simulations. Using independent Dirichlet random numbers
for posterior distributions of population allelic frequencies, we can obtain a posterior distribution of D using
Equation 1. The number of each Monte Carlo simulation is set to 10,000, so 10,000 D are calculated from the
10,000 sets of p between two populations. The posterior
probability density function is estimated on the basis of
the histogram of D with the number of classes of 100
by using the function density of S language version 4
(Chambers and Hastie 1992). For multilocus data, the
mean of the genetic distances at J loci is calculated as
j/J.
D RJj1D
Empirical Bayes procedure: The primary disadvantage of using a Bayesian analysis for allele frequency
estimation is that there is no obvious way of selecting a
reasonable prior (Lange 1995). The Dirichlet distribution with 1 k 12 is a noninformative prior
(Box and Tiao 1992). Here we adopt an empirical Bayes
procedure to avoid dependence to priors. This procedure estimates the hyperparameters by maximizing
the marginal-likelihood function (Maritz and Lwin
1989),
L(|n) P(n|p)(p|)dp

(Rki1ni)! (Rki1i)

ni! ki1(i)

k
(Rki1i)
(Rki1ni)!
(i ni)
,
(
k
n
!
(R

Rn
)
i
1
i
i
i)
i

1
i

piini1dp
i1

which is also given in Lange (1995) and Weir (1996).


The distribution is known as a Dirichlet-multinomial
distribution (Lange 1995; Weir 1996), which is a generalization of the -binomial distribution.
Lange (1995) estimated the hyperparameters from
single-locus data using Newtons method. Here we estimate them from multilocus data by maximizing Equation 3 using a simplex minimization for the negative
logarithm of Equation 3. Assuming that is the same
for H populations (samples) to be compared and Ri
is also the same for J loci, the joint marginal likelihood
is then given by
L(1j, , kj1,j(j 1, , J), 2|n)

1 ij
Cjh (Rkj 1ij i

Rkj 1nhij)
J

j1h1

(Rkj )

(ij nhij)
, (3)
(ij)
i1
kj

Empirical Bayes for Frequencies

where Cjh (Rikj 1nhij)!/ikj 1nhij! is a constant term for the


combination of the multinomial likelihood that can be
excluded from the estimation procedure.
Parameter 2 is the dispersion parameter that defines
the magnitude of overdispersion; i.e., the variance of the
response Y exceeds the nominal variance (McCullagh
and Nelder 1983). For example, the expectation of the
binomial random variables of a sample size m is E[Y]
mp and the variance is V[Y] mp(1 p). If there is
overdispersion, the variance is V[Y] 2mp(1 p)
though the expectation remains the same, where Y has
a density of a -binomial distribution. For a multinomial
event with overdispersion, the variance-covariance matrix of Y is 2 times larger than that of the multinomial
distribution, where Y has a density of a Dirichletmultinomial distribution.
Johnson and Kotz (1969) showed that the variancecovariance matrix of a Dirichlet-multinomial distribution is (Rki1ni Rki1i)/(1 Rki1i) times larger than
that of the multinomial distribution. Hence the relationship between the dispersion parameter and the hyperparameters for a population is given by
2

Rki1ni Rki1i
.
1 Rki1i

(4)

We assume equal overdispersion effects for all loci, so


the total of the hyperparameters (hereafter we use
for Ri) is the same for all loci, which gives the expression for 2 as
2n
.

1
2

(5)

Here 2n is the mean number of genes of H populations


given by 2n RhH12nh/H. Given the estimate of 2, we
have the estimator for as
2n
2
.
2

(6)

We estimate RJj1(kj 1) 1 free parameters numerically, including 2, which is assumed to be the same
among loci, and 1j, , kj1,j for locus j, and kj is
estimated by Rikj11
ij.
The binomial and multinomial counts are assumed to
be taken by a simple random sampling, so the dispersion
parameter 2 is considered to indicate the magnitude
of overshooting from a simple random sample. McCullagh and Nelder (1983) stated that The simplest and
perhaps the most common mechanism of overdispersion is clustering in the populations. Kitada et al.
(1994) estimated the dispersion parameter for fish tag
recovery data and showed that the variances of the estimated mortality rates were
2( 14.73) times larger
than those assuming the multinomial model, which was
considered to be caused by the aggregation of the
tagged fish in the fishing ground. In genetic data analysis, overdispersion corresponds to the variance of an

2065

allele frequency exceeding the nominal variance of a


simple random sample from a gamete pool. If there
are subpopulations divided spatially in a survey area, a
sample allele frequency from the area might be overdispersed even if a simple random sampling is performed. If a cluster of a genotype is taken, a sample
allele frequency from a population might be also overdispersed. One can then estimate overdispersion based
on several sets of allele frequencies obtained from the
survey area.
Standardized genetic distance: When allele frequencies at J loci are obtained from genetically identical
j/(n1. n2.) follows a 2 distripopulations, 2n1.n2. RJj1D
bution with a degree of freedom of R(kj 1) asymptotically (Nei 1987). The shape of the distribution varies
with the sample sizes and degree of freedom. When
larger numbers of individuals are sampled, the distribution is farther from 0 even if the genetic difference is
small. Suppose the case for n1. n2. n., the above
j and take a value proportional
statistics become n.RJj1D
to the sample size. It is then not convenient to make D
an index of the genetic distance.
Here we standardize D and propose a general index
for the genetic distance. Performing a square root transformation to make the variance independent of the
mean (Snedecor and Cochran 1967) and subtracting
the expected value of I (the derivation is given in the
appendix), we obtain a standardized genetic distance
as
I

j
2n1.n2.RJj1D

(n . n .)
1

((RJj1(kj 1) 1)/2)
,
(RJj1(kj 1)/2)

(7)

which follows a normal distribution with mean 0 and


variance 0.5 under the condition of p1 p2.
j/(n1. n2.) assumes
The 2 distribution of 2n1.n2.RJj1D
that 2n1 and 2n2 genes are taken by a binomial sampling
from a population. For this case, 2 in the first term of
Equation 7 equals 1. However, if there is overdispersion,
2 becomes active and takes a value larger than 1. If the
overdispersion is neglected, the genetic distance is then
overestimated and the scale of the distribution of 2n1.n2.
j/(n1. n2.) becomes 2 times larger than that
RJj1D
under the previously stated assumption. The dispersion
parameter 2 corrects this effect.
Effective population size: The effective population
size is estimated from the temporal variation of allele
frequencies in a population. Since the observed variance
of the allele frequencies includes the sampling variance
in addition to the genetic drift, we subtract the sampling
variance when estimating the effective population size.
Let us assume that we have two samples with sizes n0
and nt from the population at generations 0 and t, respectively. The empirical Bayes procedure developed
here can be extended to obtain the posterior distributions of the effective population size Ne by using the
posterior distribution of F-statistics calculated from the
posterior distribution of allele frequencies.

2066

S. Kitada, T. Hayashi and H. Kishino

The standardized variance of allele frequency change


measured by F-statistics has been used to estimate Ne
(Krimbas and Tsakas 1971; Nei and Tajima 1981; Pollak 1983; Waples 1989). Among F-statistics, Fk proposed
by Pollak (1983) is similar in form to D and is given
by
Fk

1
k1

i1

2(p0i pti)2
.
p0i pti

(8)

For the case of multiple loci, Fk is calculated by Fk


Rj(kj 1)Fkj/Rj(kj 1) from Nei and Tajima (1981),
where Fkj is Fk at the jth locus. Without overdispersion,
Ne is estimated by
e
N

2[Fk 1/(2n0) 1/(2nt) 1/N]

(9)

for plan I, where the sample is taken after reproduction.


For plan II, where the sample is taken before reproduction, the term 1/N is eliminated, where Fk is the estimator obtained by substituting sample frequencies in Equation 8 and N is the census size for a population (Waples
1989, Equations 11 and 12).
Equation 9 assumes that 2n0 and 2nt genes are taken
by a binomial sampling from the population. If there
is overdispersion, Fk is overestimated, which leads to
underestimation of Ne. Since the effective sample size
is obtained by discounting the apparent sample size by
dispersion parameters, Equation 9 is modified as
e
N

t
.
2[Fk
2/(2n0)
2/(2nt) 1/N]

(10)

CASE STUDIES

Red sea bream: To evaluate genetic distances, we first


analyzed the data of four populations of red sea bream
(Pagrus major) from Tabata and Mizuta (1997; Table
1). From the fragment pattern of mtDNA D-loop regions
with six restriction enzymes, 48 haplotypes were obtained for four wild populations. We decomposed the
haplotype frequency to six allelic frequencies for each
restriction enzyme and eliminated HaeIII and Msp,
which showed little or no polymorphism, from the analysis.
The estimate of the total hyperparameters was
106.553 for each locus (Table 1), and the dispersion
parameter was estimated at 1.80 by maximizing Equation 3. Here 2n. (72 95 93 90)/4 87.5
because mtDNA is a haploid. With a prior distribution
specified by these parameters, we obtained the posterior
distribution of D by dividing RDj over four loci by the
number of loci (Table 2). As an example, the histogram
and estimated density function of the posterior distribution of D at HinfI between Tanabe Bay and Tomogashima Channel is shown in Figure 1. D12 in Figure 2 was
obtained as the mean of such four posterior genetic

distances at the four loci. It should be noted that the


posterior distances were overestimated including the
overdispersion and the posterior distributions in Figure
2 were then overestimated. The means and SDs of D12,
D13, and D14 were about two times larger than D23, D24,
and D34; however, they might include the effect of the
smaller sample size of population 1 (Tanabe Bay).
The posterior distribution of the standardized genetic
distance took the overdispersion and sample size difference into account. The means of I12, I13, and I14 ranged
from 0.2706 to 0.4671, whereas those of I23, I24, and I34
took negative values. The SDs ranged from 0.53 to 0.58
and took almost the same values. The genetic differences with population 1(I12, I13, and I14) looked larger
than the others (Table 3). However, the posterior distributions of I overlapped well with the theoretical distribution of no genetic difference (Figure 3).
We estimated the 95% confidence interval of the dispersion parameter to be from 1.72 to 1.88 by the likelihood-ratio test. The lower limit of the dispersion parameter corresponds to the upper limit of the genetic
distance, from which we evaluate the difference. The
means of the posterior distributions for the lower limit
of the dispersion parameter were increased from 8 to
27% and SDs remained the same (Table 3), but the
posterior distributions of I were almost the same as those
for the point estimate of the overdispersion and still
overlapped well with the theoretical distribution (Figure 3).
The value of 95% upper limit of the credibility region
of the theoretical normal distribution of I with mean
of 0 and variance of 0.5 is 1.16. All posterior means were
1.16, and the credibility regions included 0; hence we
concluded that there was no genetic difference between
the four populations of red sea bream. This finding
agreed with the result of the original authors, who reported that the Roff test did not reject the homogeneity
of the haplotype frequencies (Tabata and Mizuta
1997, p 0.219). Nevertheless, they rejected the hypothesis by the pairwise comparison. From the results of our
test, however, we argue that it was inappropriate analysis.
Herring: Stock enhancement of herring (Clupea pallasii) has been performed in Akkeshi Bay, Hokkaido
(Japan). Because the matured herrings migrate to Akkeshi Bay to spawn, they are considered to have originated from Lake Akkeshi and Akkeshi Bay. Although
wild adult fish that migrated to the bay are used for
artificial spawning to produce juveniles every year, it
still may be important to monitor the genetic change
and estimate the effective population size to maintain
the wild stock.
Temporal changes in allozyme allele frequencies were
obtained by combining two studies on the same loci by
Ando and Ohkubo (1997) and Hotta et al. (1999;
Table 4). Independent samples were taken in March
and April 1993. In 1996, males and females were taken
separately from the sample, hence they were not inde-

Empirical Bayes for Frequencies

2067

TABLE 1
Allele frequencies of the mtDNA D-loop region from Tabata and Mizuta (1997) and estimated
hyperparameters for four populations of red sea bream from eastern Japan

Sample size:
HinfI

MspI
TaqI
RsaI

Tanabe
Bay
72

Tomogashima
Channel
95

Sea of
Japan
93

Bingo
Nada
90

Hyperparameter

0.458
0.375
0.167
0.903
0.097
0.958
0.042
0.097
0.264
0.306
0.180
0.153

0.411
0.442
0.147
0.863
0.137
0.811
0.189
0.137
0.305
0.316
0.116
0.126

0.376
0.409
0.215
0.882
0.118
0.806
0.194
0.215
0.312
0.312
0.075
0.086

0.378
0.456
0.166
0.867
0.133
0.856
0.144
0.178
0.266
0.289
0.078
0.189

43.855
43.872
18.826
92.864
13.688
91.248
15.305
17.015
30.253
32.871
11.802
14.613

pendent. For the purposes of our case study, we treated


the data as if they were taken independently.
The estimate of the total hyperparameters was
130.956 for each locus (Table 4), and the dispersion
parameter was estimated at 2.39, with 2n. (206
214 168 148)/4 184. The posterior distribution
of Fk estimate was calculated from each of two sets of
posterior distributions of allele frequencies for 1993 and
1996 by
1 RJ (k 1)Fkj
.
(11)
Fk jJ 1 j
4
Rj1(kj 1)

tion, so Equation 10 was used by eliminating the term


of 1/N and substituting (206 214)/2 210 for 2n0
and (168 148)/2 158 for 2nt. The posterior means
of Ne estimates and 95% credibility region of Ne are
given in Table 5.
The dispersion parameter was estimated at 2.39 with
a 95% confidence interval from 1.00 to 7.51. From a
conservation viewpoint, it is better to consider the lower
limit of Ne. The lower limit of the dispersion parameter
evaluates the upper limit of Fk corresponding to the
lower limit of Ne. The lower limit of 2 was 1.00. Corresponding with that, no overdispersion arose and no
subpopulation existed. The number of simulations with
a negative value of Ne estimate was 1221 in 10,000 trials.
When Fk [1/(2n0) 1/(2nt)], the only feasible estimate of Ne is infinity (Waples 1989). The mean of the
positive Ne estimate was 350, and the 95% credibility
region was estimated from 20 to infinity (Table 5). The
posterior distribution of the positive Ne estimate is
shown in the right side of Figure 4; it suggests the order
of the effective population size of herring, even though
the upper limit was not estimated.
Northern pike: We analyzed the data from Miller
and Kapuscinski (1997) for comparison. Temporal
changes in microsatellite allele frequency at seven loci
were typed in the northern pike (Esox lucius) population
of Lake Escanaba, Wisconsin. Of the seven loci, five had

Fk ranged from 0.0014 to 0.0814 with the mean and


SD of 0.0226 and 0.0105, respectively. The posterior
distribution of Fk is shown in the left side of Figure 4.
Most of the matured herring migrating to Akkeshi
Bay to spawn are in their second year of life; the remainder are in their third year. The age composition of the
spawners was surveyed and estimated at 0.9 and 0.1 for
each age class by the Japan Sea-Farming Association.
The expected number of generations can be used for
t in the estimating equations of Ne because the expectation of the F-statistics was approximated to be linear
with t as shown in Waples (1989, p. 382). We estimated
the expected number of generations at 1.48 ( 3/2.1)
divided by the number of years between samples by
the mean age of spawners as Miller and Kapuscinski
(1997) did. The samples were taken before reproduc-

TABLE 2
Means and SDs of the posterior distribution of the genetic distance for the red sea bream populations

Mean
SD

D12

D13

D14

D23

D24

D34

0.0585
0.0207

0.0523
0.0196

0.0530
0.0192

0.0229
0.0112

0.0241
0.0116

0.0244
0.0118

2068

S. Kitada, T. Hayashi and H. Kishino

Figure 1.A histogram and the estimated density function


of the posterior distribution of the genetic distance between
red sea bream populations of Tanabe Bay and Tomogashima
Channel for HinfI using data given in Table 1. D12 in Figure
2 was obtained as the mean of the four posterior genetic
distances for the four loci.

two alleles and two had three alleles. Following the data
processing techniques of Williamson and Slatkin
(1999), we also combined the two least common allelic
classes for the two loci with three alleles and created a
diallelic frequency set. To estimate the hyperparameters, we allocated sample sizes of 86 for 1961, 110 for
1977, and 72 for 1993 according to the diallelic frequencies for each locus to obtain the number of individuals
corresponding to the frequencies. Because the data had
one sample for each year, it was not possible to estimate
the hyperparameters for each locus. Therefore, we assumed that the seven loci were independent of each
other and had common hyperparameters. The two common hyperparameters and the dispersion parameter
were estimated at 16.232, 3.089, and 9.74, with 2n.
(172 220 144)/3 178.7.
Fk between the year of 1977 and 1993 ranged from
0.0087 to 0.1264 with the mean and SD being 0.0479
and 0.0150, respectively. The posterior distribution of
Fk is given in the left side of Figure 5. The posterior
distribution of the Ne estimate was obtained by substituting the posterior distribution of Fk into Equation 10,
eliminating the term 1/N, with t 4, as given in Miller

Figure 2.Posterior distributions of the genetic distance


(D) between four populations of red sea bream using data
given in Table 1.

and Kapuscinski (1997). The mean of the positive Ne


estimate was 73 and a 95% credibility region neglecting
the overdispersion was estimated from 29 to 190 (Table
6). The number of negative Ne estimates was 6 in 10,000
trials. Miller and Kapuscinskis estimates for 1977 and
1993 data were 0.038 for Fk and 72 for Ne with a 95%
confidence interval from 17 to 258. Our posterior mean
of Fk was 1.26 times larger and that of the Ne estimate
coincided with a 67% narrower confidence interval (Table 6).
The 95% confidence interval of the dispersion parameter was estimated from 5.51 to 18.80. For the 95%
lower limit of the dispersion parameter, the number of
simulations with a negative value of Ne estimate was 8480
in 10,000 trials. The mean of the positive Ne estimates
was 1065 and the 95% credibility region was estimated
from 123 to infinity (Table 6). The posterior distribution of the positive Ne estimate, neglecting the overdispersion (2 1), and for the lower limit of
2(
5.51) are given in the right side of Figure 5. In this
example, we can see the effect of the overdispersion on
the posterior distribution of Ne estimate. The estimate
of Ne, neglecting the overdispersion, agreed well with
the estimate of Miller and Kapuscinski (1997).
Ayu broodstock: Ayu (Plecoglossus altivelis) is the most
popular target species of recreational anglers in rivers
and streams in Japan. A total of 300 million juveniles

Empirical Bayes for Frequencies

2069

TABLE 3
Means and 95% credibility regions of the posterior distribution of the standardized genetic distance
for the red sea bream populations
2 1.80a

I12
I13
I14
I23
I24
I34

Mean

SD

0.4671
0.2706
0.2735
0.6241
0.5866
0.5768

0.5774
0.5747
0.5567
0.5284
0.5332
0.5276

2 1.72b
95% CR

[0.4819,
[0.6706,
[0.6259,
[1.4582,
[1.4226,
[1.4047,

1.4232]
1.2169]
1.1977]
0.2767]
0.3163]
0.3319]

Mean

SD

0.5448
0.3435
0.3464
0.5729
0.5345
0.5244

0.5914
0.5886
0.5702
0.5411
0.5461
0.5404

95% CR
[0.4272,
[0.6205,
[0.5747,
[1.4272,
[1.3907,
[1.3723,

1.5240]
1.3127]
1.2931]
0.3497]
0.3903]
0.4062]

CR, credibility region.


a
Point estimate of the dispersion parameter.
b
95% lower limit of the dispersion parameter.

are released every year, of which hatchery-produced fish


comprise 30%. The life span of ayu is 1 year. They
spawn in a river from September to November and die
after spawning. Hatched larvae go down to the sea and
winter there. The upstream run of wild ayu juveniles
begins from the coast in late March to early April and
is over by early July. Soon after, they mature, spawn from
September to November, and then die after spawning.
Hatcheries have commonly cultured broodstocks over
generations. At the Gunma Prefecture Fisheries Experimental Station, adult ayu have been cultured over 27
generations. About 30004000 fish have been reared
every year as broodstock, some of which are used for

artificial fertilization. In 1996, 850 females and 650


males were used. The temporal changes in allozyme
allele frequencies of the ayu broodstock were reported
by Yoshizawa (1997; Table 7). These samples in Table
7 were taken after artificial fertilization from two rearing
tanks in which males and females were kept separately.
The total of the hyperparameters was estimated at
28.576 for each locus (Table 7), and the dispersion
parameter was estimated at 5.86, with 2n. (190
210 128 130 100 100)/6 143. The 95%
confidence interval of the dispersion parameter was estimated to range from 3.04 to 13.49. We calculated four
Fks on the basis of temporal changes in allele frequen-

Figure 3.Posterior distributions of the standardized genetic distance (I ) for the red sea bream populations taking the point
estimate of the dispersion parameter 1.80 (left) and the 95% lower limit 1.72 (right) into account.

2070

S. Kitada, T. Hayashi and H. Kishino


TABLE 4
Temporal changes in allozyme allele frequencies and estimated hyperparameters
of herring in Akkeshi Bay
1993a
Sample size:
Gpi
Pgm
a
b

1996b

March
103

April
107

Male
84

Female
74

Hyperparameter

0.282
0.718
0.447
0.553

0.243
0.757
0.393
0.607

0.191
0.809
0.321
0.679

0.149
0.851
0.324
0.676

28.630
102.326
48.993
81.963

Ando and Ohkubo (1997).


Hotta et al. (1999).

cies observed in the first (19961997; F1) and second


(19971998) time intervals (F2), over the entire interval
(19961998; F3), and for the entire interval based on
the pooled F for the first two intervals, as Miller and
Kapuscinski (1997) did. For the last case, Miller and
Kapuscinski (1997) used the sum of F1 and F2 for the
entire interval, but we used the mean for the two intervals (Fmean), which gives the same form of Ne estimate
by Equation 10, substituting n0 by the harmonic mean
of the sample size of the first and second year, and nt
by that of the second and third year. When t is equal
for the two intervals, the estimate of Ne derived from
pooled Fk is the harmonic mean for both sampling periods (Nei and Tajima 1981; Pollak 1983; Waples 1989;

Miller and Kapuscinski 1997; Williamson and Slatkin 1999).


The values of Fk calculated by Equation 11 were varied
using four combinations for two samples in each sampling year, reflecting the variation in allele frequencies
for each sampling period. The posterior mean and SD
of F3 were the largest, and the SD of Fmean was the smallest
but almost the same as that of F1 (Table 8). The posterior distributions of Fk are illustrated in the left side of
Figure 6.
We fixed the generation time at t 1.0 for the first
and second time intervals and had t 2.0 for the entire
interval because the life span of ayu is 1 year. The samples were taken after reproduction, so we used Equation

Figure 4.Posterior distributions of the standardized variance of allele frequency change Fk and effective population size Ne
of herring for the 95% lower limit of the dispersion parameter (2 1) using data in Table 4.

Empirical Bayes for Frequencies

2071

TABLE 5

TABLE 6

Means and 95% credibility regions of the posterior


distribution of the effective population size of herring
in Akkeshi Bay obtained using data in Table 4

Means and 95% credibility regions of the posterior


distribution of the effective population size of
northern pike obtained using 19771993 data
from Miller and Kapuscinski (1997)

Meana

95% CR

/10,000b

1.00c
2.39d
7.51e

350
16,841

[20]
[35]
[]

1,221
6,832
10,000

Mean of positive Ne.


Number of Ne that took in 10,000 simulations.
c
95% lower limit of the dispersion parameter (no overdispersion).
d
Point estimate of the dispersion parameter.
e
95% upper limit of the dispersion parameter.
b

10, substituting 2n0 (190 210)/2 200, 2nt


(100 100)/2 100, and N 1500, which was the total
number of individuals used for artificial fertilization.
We made four estimates of Ne on the basis of F1, F2,
F3, and Fmean. Estimates for the 95% lower limit of the
dispersion parameter ( 3.04) are given in Table 8.
The posterior mean obtained using F3 was the largest
with the largest SD, and that for Fmean was second with
a smaller SD. The Ne estimate that took in 10,000
simulations was 8677 when using Fmean, and that for F3
was 4927. This is because the sampling correction term
in the denominator of Equation 10 took the very similar
values of 0.0912 for F3 and 0.0928 for Fmean, despite the

Meana

95% CR

1.00c
M and Kd
5.51e
9.74f

73
72
1,065
20,606

[29190]
[17258]
[123]
[]

/10,000b
6
8,480
9,998

Mean of positive Ne.


Number of Ne that took in 10,000 simulations.
c
No overdispersion is assumed.
d
Estimated by Miller and Kapuscinski (1997).
e
95% lower limit of the dispersion parameter.
f
Point estimate of the dispersion parameter.
b

posterior mean of Fmean being smaller than that of F3,


as shown in the left side of Figure 6. The posterior
distributions of the positive Ne estimate obtained by
using F1, F2, F3, and Fmean for the lower limit of
2 are
given in the right side of Figure 6, showing a larger
variance of the estimate based on F3 than those derived
from F1, F2, and Fmean.
We failed to estimate the upper limit of the credibility
regions because of large sampling variances with overdispersion. However, when the numbers of breeding
males Nm and females Nf are given, which is difficult to

Figure 5.Posterior distributions of the standardized variance of allele frequency change Fk and effective population size Ne
of northern pike for the 95% lower limit of the dispersion parameter (2 5.51) and that with no overdispersion (2 1.00)
using 19771993 data from Miller and Kapuscinski (1997).

2072

S. Kitada, T. Hayashi and H. Kishino


TABLE 7
Temporal changes in allozyme allele frequencies from Yoshizawa (1997) and estimated hyperparameters of
ayu broodstock cultured over 27 generations in Gunma Prefecture
1996
Sample size:
Gpi1
Mpi1

1997

1997

Male
95

Female
105

Male
64

Female
65

Male
50

Female
50

Hyperparameter

0.184
0.816
0.521
0.479

0.200
0.800
0.400
0.600

0.141
0.859
0.422
0.578

0.192
0.808
0.315
0.685

0.290
0.710
0.280
0.720

0.230
0.770
0.300
0.700

6.341
22.235
10.794
17.783

know in a wild population but possible in hatcheries,


the effective population size is obtained by Ne 4NmNf/
(Nm Nf) (Wright 1931). We obtained a value for Ne
of 1473 by using the equation (4 850 650/(850
650)), which referred to the effective population size
where a random mating was performed by artificial fertilization. In a spawning season of ayu, males and females eligible for spawning were selected every day from
the broodstock and used for artificial fertilization. The
number of females used in an artificial fertilization
ranged from 10 to 20, and the ratio of males to females
was 0.8. The eggs were squeezed from the females
and stocked in a stainless bowl and then fertilized by
squeezing sperm from individual males. This method
of fertilization might not guarantee a random mating
of the males and females used; hence, 1473 should be
used as the upper limit of the credibility regions instead
of (Table 8). If we neglect overdispersion, the 90%
credibility region could be obtained at [13589] with
the posterior mean of 136, which was underestimated.
DISCUSSION

The empirical Bayes procedure developed here


makes it possible to describe the skewness of the genetic
distance and evaluate genetic differentiation between

populations while taking full account of the uncertainty


of the sample allele frequencies. When we compare
populations in which the genetic differentiation is small,
the hypothesis-testing framework cannot accept the null
hypothesis of no genetic differentiation in almost all
cases, because of the poor statistical power with relatively
small sample sizes. The empirical Bayes procedure is
effective even in such cases. So we believe it could play
an important role in the field of conservation.
This general method can easily be extended to any
parameter that is a function of multinomial frequencies.
When the parameter of interest is a function of allele
frequencies, the posterior distribution of that parameter
can be obtained through the function by using the posterior distribution of the allele frequencies, instead of
assuming a prior distribution for the parameter.
Overdispersion and empirical Bayes: Until now, models based on a simple random sampling from the gamete
pool have been assumed when evaluating allele frequencies. However, as shown in the four case studies treated
in this article, a simple random sampling is not necessarily guaranteed. If there are subpopulations divided spatially in a survey area, or a cluster of a genotype is taken,
a sample allele frequency might be overdispersed. If
overdispersion arises, a sampling variance becomes 2
times larger than that for a simple random sampling.

TABLE 8
Means and 95% credibility regions of the posterior distribution of the effective population size of ayu for
95% lower limit of the dispersion parameter (3.04) obtained using data in Table 7
Fk
Period
19961997
19971998
19971999
19971999
19971999
a

t
(F1)
(F2)
(F3)
(Fmean)
(Fmean)d

1
1
2
1

Mean

0.0263
0.0379
0.0474
0.0321

Ne
SD

Mean

SD

95% CR

/10,000c

0.0121
0.0189
0.0191
0.0120

350
240
796
491
136

4,024
1,388
18,538
3,294
3,720

[32]
[17]
[22]
[35]
[13589e]

8,453
8,094
4,927
8,677
377

Number of generations.
Mean of positive Ne.
c
Number of Ne that took in 10,000 simulations.
d
No overdispersion is assumed (2 1).
e
90% credibility region.
b

Empirical Bayes for Frequencies

2073

Figure 6.Posterior distributions for the four estimators of the standardized variance of allele frequency change Fk and
effective population size Ne of ayu broodstock using 19961998 data in Table 7.

This can seriously affect the precision of the estimate


of genetic distance and the effective population size.
As a result, the genetic distance and F-statistics can be
overestimated, and the effective population size can be
underestimated, if overdispersion is not taken into account in the analysis. Therefore, it is quite important
to take overdispersion into account when estimating
genetic distance and effective population size.
Sample sizes: If we use the noninformative prior of
the Dirichlet distribution (Box and Tiao 1992), the
dispersion parameter might be overestimated for most
cases. For example, the dispersion parameter of herring
was estimated at 92.5 from the noninformative prior,
which was estimated at 2.39 from the empirical Bayes
procedure, illustrating the importance of estimating the
hyperparameters from the data.
We examined the relationship between sample size
and the estimates of the hyperparameters using the data
of red sea bream given in Table 1. We estimated the
hyperparameters with multipliers of 0.5, 2, 3, and 4
to test each population with the same sample allele
frequencies. The estimates of the hyperparameters were
stable and not dependent upon sample sizes. This confirmed the robustness of the empirical Bayes procedure
(Table 9). However, the dispersion parameter became
larger as sample size increased. This is to be expected
from the relationship between the total of the hyperparameters, the sample sizes, and the dispersion parameter
given by Equation 5.
Suppose there are several subpopulations and the
population allele frequencies are largely varied spatially.

If the sample sizes are small, one might consider that


the large variation in the sample allele frequencies is a
function of the small sample size. On the other hand,
if the same allele frequencies are obtained for larger
sample sizes, one could consider that the large variation
comes from the subpopulation structure with confidence. The more samples one draws, the more precisely
one can estimate the dispersion parameter. In addition,
increasing the number of polymorphic loci to be surveyed may also increase the information available for
estimating the dispersion parameter, e.g., the precision
of the dispersion parameter estimate in the case of red
sea bream, in which the narrowest confidence interval
was obtained among our four case studies.
It is also quite important to consider sampling strategies to minimize overdispersion caused by sampling procedures. For example, sampling from different sites and
times may be useful to avoid sampling clusters of individuals having the same genotype. Such multiple samples
contribute simultaneously to a more precise estimation
of the dispersion parameter.
Standardized genetic distance: As I follows the normal distribution, it takes values between and . For sim j/((n1. n2.)2) and define the
plicity, let X 2n1.n2.RJj1D
expected value by E[X]. If X E[X], I takes a positive value, and 0 if X E[X]. If X E[X], I takes
a negative value. As X follows the 2 distribution asymptotically when there is no genetic difference, E[X] is
almost equal to the square root of the degrees of freedom of X, which is the number of loci examined. When
j does not increase compared to the increased numRJj1D

2074

S. Kitada, T. Hayashi and H. Kishino


TABLE 9
Estimated hyperparameters and the dispersion parameter for four populations of red sea bream for sample
sizes of 0.5, 2, 3, and 4 times larger than the original one with the same sample allele frequencies
Sample size:

2
HinfI

MspI
TaqI
RsaI

0.5

Original

106.844
1.40
43.956
43.989
18.899
93.049
13.795
90.981
15.863
16.961
30.155
32.976
11.841
14.910

106.553
1.80
43.855
43.872
18.826
92.864
13.688
91.248
15.305
17.015
30.253
32.871
11.802
14.613

106.168
2.62
43.603
43.887
18.679
93.007
13.161
90.853
15.315
16.759
30.091
33.032
11.777
14.509

106.119
3.44
44.301
43.932
17.886
92.027
14.093
91.949
14.170
16.886
29.762
32.855
11.732
14.885

106.075
4.25
43.683
43.986
18.406
93.033
13.042
91.253
14.822
16.383
30.310
32.791
11.910
14.682

ber of loci examined, the likelihood for I taking a negaj


tive value increases. On the other hand, when RJj1D
increases to the increased number of loci examined,
the likelihood for I taking a positive value increases.
This point illustrates the effectiveness of increasing the
number of loci to obtain increased information on the
genetic differentiation from the value of the posterior
mean of I. Conversely, a negative posterior mean indicates that little information on genetic differentiation
will be obtained even if the number of loci is increased,
as a function of the small genetic differentiation. This
is considered to be the cause of the negative values of
the posterior mean for I23, I24, and I34.
Overdispersion and gene flow: Weir (1996) stated that
for populations that have reached an equilibrium under
the joint effects of drift and mutation or migration,
Wright (1945) found that allele frequencies for loci
with two alleles had a distribution, and for multiallele
loci the distribution was Dirichlet (Wright 1951). We
assumed that the hyperparameters for the or Dirichlet
distributions were common for every sample and locus
that was in an overdispersed population. Our assumption corresponds with Wrights (1945, 1951) theories.
So if the random sampling is performed, the estimated
hyperparameters and dispersion parameter both describe a kind of genetic differentiation between populations that have reached an equilibrium. If all populations mate randomly, the total variance of allele
frequency p with two alleles of 2n genes is given by Weir
(1996)
V(p)

p(1 p)
{FST(2n 1) 1},
2n

(12)

2 FST(2n 1) 1,

from which we can see larger FST gives larger overdispersion. From Equation 13, we also have the relationship
FST

2 1
.
2n 1

(14)

Rannala and Hartigan (1996) proposed the pseudomaximum-likelihood method (PMLE) for estimating
the rate of gene flow into island populations using the
distribution of alleles in samples from a number of
islands. We confirmed that their likelihood function
for multiple loci (p. 149 Equation 10) coincides with
Equation 3 by using the relationship of (n)
(n 1)!. In PMLE, i is treated by pi. Here, pi is a
TABLE 10
Estimated hyperparameters and the dispersion parameter
from the mtDNA haplotype distribution among islands
for Channel Island foxes (Table 2 of Rannala
and Hartigan 1996), using the full-likelihood
function (Equation 3)

Parameter

Rannala and
Hartigan (PMLE)
0.41
(0.35)a
0.1189
0.0574
0.0082
0.0779
0.1476
18.38c

1
2
3
4
5
2
a

where FST is the coancestry coefficient of Wright


(1951). The second term of Equation 12 corresponds
to the dispersion parameter, yielding the relationship

(13)

This article
(MLE)
0.45
[0.18, 0.84]b
0.0945
0.0428
0.0382
0.0992
0.1760
17.89

SD.
95% confidence interval estimated by the likelihood-ratio
test.
c
Estimated by Equation 5.
b

Empirical Bayes for Frequencies

nuisance parameter estimated from the data as pi ni./


n . ., and then pi is substituted for pi in the log-likelihood
function, and the only unknown parameter is estimated by using the Newton method. By contrast, we
directly maximize the negative log-likelihood function
and estimate RJj1(kj 1) 1 parameters by using a
simplex minimization. We estimated the hyperparameters and the dispersion parameter from the mtDNA haplotype distribution among islands for Channel Island
foxes given in Table 2 of Rannala and Hartigan
(1996) using the full-likelihood function (Equation 3).
The estimates were similar to Rannala and Hartigans
PMLE (Table 10). Thus, our empirical Bayes procedure
also offers the maximum-likelihood estimators (MLEs)
of the rate of gene flow. MLE is more efficient than
PMLE (Chuang and Cox 1985) and has the advantage
that it can estimate the confidence interval of the parameters by using the likelihood-ratio test.
Wright (1969) proposed the estimator of for a
discrete-generation island model of a population at
equilibrium, based on FST as 1/FST 1 (Rannala
and Hartigan 1996). Substituting this estimator into
Equation 4, we have Equation 13, which was obtained
from the total variance of Weir (1996). As is clear from
Equation 6, larger 2 gives smaller , indicating that
larger genetic differentiation corresponds to smaller
gene flow. For the case of the red sea bream, 2 and
were estimated at 1.80 and 106.55, respectively. For the
case of the foxes, they were estimated at 17.89 and 0.45,
respectively. From this result, it is clear that the six fox
populations in the isolated islands had small gene flow
and large genetic differentiation. On the other hand,
red sea bream had large gene flow and small genetic
differentiation. The estimate of FST for red sea bream
was FST (1.80 1)/(87.5 1) 0.0093, which was
relatively small. But for foxes it was FST (17.89
1)/(25.5 1) 0.6894, suggesting advanced inbreeding in the fox populations.
The essential idea for estimating overdispersion is
to compare the variation of sample allele frequencies
obtained from the different locations to the multinomial variance. In addition, the effective population size
is based on the changes in allele frequencies between
generations. Conversely, overdispersion provides insight into the spatial variation of allele frequencies. By
evaluating the spatial variation, it might become possible to discriminate the overdispersion resulting from
the variation between generations. Hence, the procedure needs to evaluate overdispersion as a function of
the spatial variation and then measure the variation
between generations taking overdispersion into account.
In the three case studies we looked at for estimating
the effective population size, direct information on the
spatial variation was scarce. Therefore, the precision of
the dispersion parameter was marginal. When subpopu-

2075

lations exist, overdispersion arises and affects the estimation of the effective population size. It is then important
to collect data on the spatial variation. At the same time,
when many isolated subpopulations exist, the effective
population size is considered to be close to the size of
a subpopulation. When this occurs, it seems dangerous
to dismiss the variation between generations as overdispersion. It needs further consideration.
Practical considerations on estimating Ne: From the
approximate variance formula of Ne estimate (Pollak
1983, Equations 28 and 29; Waples 1989, Equation 17),
it is clear that increasing the sample size, the number
of loci, and the number of generations t simultaneously
ensures greater precision for the estimate of Ne (Waples
1989). Miller and Kapuscinski (1997) stated that if
Ne is expected to be moderately large, the sample size,
the number of loci, and the number of generations
should all be as large as possible. To improve the precision of the estimate of Ne, it is essential to reduce the
sampling variance and increase information on genetic
drift.
Sample size: The idea of the temporal method is to
estimate Ne from the genetic change over time described
by F-statistics estimated from the sample allele frequencies. F-statistics, then, consist of the genetic drift and
the sampling variance. To evaluate the genetic drift, we
have to subtract the sampling variances from the
F-statistics. The second and third terms in the denominator of Equation 10 are the sampling variances at generations 0 and t. If Ne is large, the genetic drift may be
small, so the denominator of Equation 10 would take
a negative value, which leads to an infinite Ne for small
sample sizes n0 and nt. If overdispersion arises, the effect
of subtracting the sampling variances becomes 2 times
larger, which is why we failed to estimate the upper limit
of the credibility region of Ne. As pointed out by Waples
(1989), the temporal method should be useful for cases
of small Ne, where larger genetic drift is expected. Even
in the case of a small Ne, the problem of an infinite Ne
estimate can occur due to large sampling variance, as
shown in the ayu studies, because of the small sample
sizes. When one uses the temporal method, reducing
the sampling variance is indispensable. The sample size
should be kept as large as possible. A larger sample size
also provides greater information on the genetic drift.
Number of loci: Williamson and Slatkin (1999) developed a maximum-likelihood temporal method to estimate Ne and compared estimates with those derived with
the F-statistic method. The simulation result in their
Table 1 showed that increasing the number of loci reduced the variance and bias in both estimators, although
when the number of loci was 50, the corresponding
reduction of variance and bias was not large, and the
total information on allele frequency changes did not
increase much. The results of Williamson and Slatkin
(1999) suggest that information on genetic drift was
not improved much even if the number of loci was

2076

S. Kitada, T. Hayashi and H. Kishino

100, because their simulation was based on diallelic


alleles. F-statistics measure a magnitude of changes in
allele frequencies per allele, which can be regarded as
a sample mean. So, the estimation precision can be
improved if the number of loci is increased. This suggests that increasing the number of alleles is essential,
which can be attained by increasing the number of loci.
Number of years between samples: The number of years
between samples is correlated with the number of generations, and it then affects the precision of the estimate
of Ne. A large number of generations between samples
can improve the precision of the estimate of Ne (Waples
1991), because information on genetic drift increases
as the number of generations increases. Williamson
and Slatkin (1999, Table 1) showed through simulated
populations sampled at generations 04 and 08 that
the variance and bias in both estimators were reduced
when the number of years between samples was doubled, although the effect of reducing the bias was not
clearly observed with the F-statistic method.
For the case study of ayu, the posterior mean of Ne
was 350 based on F1 and 796 based on F3, and SDs for
the two estimates were 4024 and 18,538, respectively,
showing the reduction of precision despite the fact that
the number of years between samples was doubled (Table 8). This is because the doubled number of generations increased the variance of allele frequency changes.
The numbers of infinite Ne estimates in 10,000 simulations were 8453 on F1 and 4927 on F3, and the smaller
F values increased the estimated value of Ne. The result
was similar for northern pike. The point estimate of Ne
based on F3 ( 125) was larger than those based on F1
( 35) and F2 ( 72), and the confidence interval for
F1 was the largest (Miller and Kapuscinski 1997).
Miller and Kapuscinski (1997) discussed the question of which estimate obtained from F3 and Fmean to use
for the entire time interval. If Ne changes between the
two sampling intervals, it should be evaluated by using
F1 and F2 for the respective intervals. The numbers of
adult ayu used for the artificial fertilization in 1997 were
600 females and 480 males, which are lower numbers
than those used in 1996. The posterior mean of F2 was
larger than that of F1, which resulted in a smaller Ne
estimate based on F2, supporting the fact that Ne actually
changed (Table 8 and Figure 6). One can use the estimate of Ne based on Fmean as the harmonic mean of the
effective population size in the respective intervals. The
precision was improved by using Fmean, in which a larger
quantity of information was included. The greater precision that occurred with using Fmean and the lower precision that occurred with using F3 for the case study of
ayu are shown clearly in Figure 6.
When Ne does not change in the entire interval, F3 is
expected to have more information on genetic drift
than F1 and F2. Williamson and Slatkin (1999, Table
1) also showed by their simulation that the estimates of
Ne based on Fmean had smaller variances and biases than

those based on F1 and F3. This suggests that the decision


about which estimate obtained from F3 and Fmean to use
should be made on the basis of the relative effect of
improving precision by using Fmean and a doubled number of years. Which estimate has more information on
the genetic drift must be determined on a case-by-case
basis.
Overlapping generations: The basic theory of the temporal method assumes generations to be discrete. The
expected number of generations used in Equation 10
directly affects the estimate of Ne. We take time to be
measured in years. The expected number of generations
between samples can be estimated by dividing the number of years between samples by the mean generation
time, which corresponds to the mean age of maturity.
In the case of ayu, since the life span is 1 year, 1 year
coincides with one generation, which makes it possible
to estimate E[t] by the above-mentioned method.
In the case of herring and salmon, however, where
there are overlapping year classes of spawners, the estimation of E[t] is complex. When generations overlap,
the age-specific birth rate may essentially affect the estimate of E[t]. Hill (1979) showed that estimates of Ne
are robust with overlapping generations if demographic
parameters are stable. If demographic parameters
change over time, F may be biased upward, leading to
an estimate of Ne that is too small (Pollak 1983; Waples
1989). Jorde and Ryman (1995) proposed a correction
method for the bias and showed by using simulations
for short time intervals that the bias was larger for a case
where age-specific birth rates were extremely different
compared with a case with equal age-specific birth rates.
Waples (1990) developed a statistical method for this
situation that can be applied to Pacific salmon populations that have an unusual life history of semelparity
with overlapping year classes. Tajima (1992) developed
a formula to estimate the expected number of generations without computer simulations and showed the estimates agreed well with estimates obtained by the
method of Waples (1990), which requires computer
simulations.
In the case of herring, age-specific survival and birth
rates were unknown, so it was not possible to apply the
method of Jorde and Ryman (1995), which requires
these demographic parameters. If an individual continues to spawn every year after the first spawning, like
herring, E[t] may be estimated by dividing the years
between samples by the mean age of spawners, leading
to the estimate of E[t] at 1.48 ( 3/2.1). If a distribution
of age-specific birth rates is concentrated to a specific
age, E[t] may be close to the estimate obtained by the
methods of Waples (1990) or Tajima (1992). We estimated the number of steps at 12 and the expected
number of generations at 5.71 ( 12/2.1) for a time
interval of 3 yr between samples by using the computer
program given in Tajima (1992), where we substituted
f(2) 0.9, f(3) 0.1, and f(4) 0. The downward bias

Empirical Bayes for Frequencies

Figure 7.Posterior distributions of the rate of inbreeding


of herring, ayu broodstock, and northern pike for the 95%
lower limit of the dispersion parameter.

e when demographic parameters change over time


of N
with overlapping generations should be corrected upward. From a conservation viewpoint, the estimate of
Ne without the correction must be conservative for the
overlapping generations.
Rate of inbreeding: As another evaluation of breeding
population size, the inbreeding coefficient may be useful, especially for cases where the population size is
estimable, as it is in the field of fishery science. Crow
(1954) used the inbreeding effective size, making a distinction between that and the variance effective size,
which was defined by an inverse of the inbreeding coefficient. However, it is known that there is no great difference between the two effective sizes (Nei 1987); hence
we calculated the posterior distribution of the rate of
inbreeding defined as 1/(2Ne) (Falconer and Mackay
1996), as an infinite Ne corresponds to an inbreeding
coefficient of 0, obtained from the 95% credibility region. The means, SDs, and 95% credibility regions of
the posterior distributions of the rate of inbreeding
were 0.0083, 0.0070, and [0, 0.0251] for herring; 0.0004,
0.0012, and [0, 0.0041] for northern pike; and 0.0011,
0.0041, and [0, 0.0140] for ayu.
The posterior mean of herring was 7.5 times larger
than that of ayu with a right-tailed credibility region
shown in Figure 7. Overdispersion for herring was underestimated, because the samples for males and fe-

2077

males taken in 1996 were analyzed as independent samples even though they were the same sample, leading
to a smaller value of Ne, which caused the rate of inbreeding to be overestimated. The mean for northern pike
was smallest with the narrowest credibility region. However, these values may be underestimated because of an
overestimated dispersion parameter of northern pike,
which was the largest among our four case studies. There
was only one sample for one sampling year, and we
assumed that the seven loci had common hyperparameters, so the estimated dispersion parameter may include
the change of the allele frequencies.
Multistage sampling in hatcheries: All existing methods assume that Ne is drawn from a gamete pool by a
simple random sampling. This is an appropriate assumption for the reproduction of a wild population. However,
for broodstocks cultured over generations in hatcheries,
candidates of the next broodstock are sampled from
the progenies produced by the broodstock. Therefore,
Ne is drawn from the progenies by a two-stage sampling.
If artificial fertilization using a part of the candidates is
performed, as in the case study of ayu, Ne is drawn from
the progenies by a three-stage sampling and the sample
is drawn from the candidates to estimate the allele frequencies, which is therefore a two-stage sampling of the
progenies. The multistage sampling must lead to the
different form of V(x y) given in Waples (1989). This
is a problem that needs further research, but it should
be noted that the variances corresponding to the twostage and three-stage sampling become small when the
sample sizes are large. In the case of ayu, a total of
30004000 candidates were sampled from the progenies
and cultured in rearing tanks, and 1500 adult fish from
the candidates were used for artificial fertilization.
Hence the sample allele frequencies of ayu were expected to represent those of the progenies produced
by the broodstock. However, if the sample sizes are
small, V(x y) is seriously affected.
We thank Zhao-Bang Zeng and two anonymous referees for their
comments on an earlier version of this article. We also thank Ray Timm
for critical review of the manuscript, Fumio Tajima for important
suggestions made during our research, Kazutomo Yoshizawa for biological information on ayu broodstocks including unpublished data,
and Masashi Yokota for helpful discussions.

LITERATURE CITED
Ando, T., and N. Ohkubo, 1997 A study for estimating effective
population size from allozyme allele frequencies, Heisei 8 nendo
seitaikei hozengata syubyo seisan gijyutu kaihatsu kenkyu seika no gaiyo,
pp. 7479. Fishery Agency of Japan, Tokyo (in Japanese).
Bartley, D. M., 1999 Marine ranching, a global perspective, pp.
7990 in Stock Enhancement and Sea Ranching, edited by B. R.
Howell, E. Moksness and T. Svasand. Blackwell, Oxford.
Bartley, D. M., D. B. Kent and M. A. Drawbridge, 1995 Conservation of genetic diversity in a white seabass hatchery enhancement
program in southern California. Am. Fish. Soc. Symp. 15: 249
258.
Blankenship, H. L., and K. M. Leber, 1995 A responsible approach
to marine stock enhancement. Am. Fish. Soc. Symp. 15: 165175.

2078

S. Kitada, T. Hayashi and H. Kishino

Box, G. E. P., and G. C. Tiao, 1992 Bayesian Inference in Statistical


Analysis. Wiley, New York.
Campton, D. E., 1995 Genetic effects of hatchery fish on wild populations of Pacific salmon and steel head: what do we really know?
Am. Fish. Soc. Symp. 15: 337353.
Chambers, J. M., and T. J. Hastie (Editors), 1992 Statistical Models
in S. Wadsworth and Brooks, Belmont, CA.
Chuang, C., and C. Cox, 1985 Pseudo maximum likelihood estimation for the Dirichlet-multinomial distribution. Comm. Stat. Theory Methodol. 14: 22932311.
Cohen, J., 1994 The earth is round (p 0.05). Am. Psychol. 49:
9971003.
Crow, J. F., 1954 Breeding structure of populations. II. Effective
population number, pp. 543556 in Statistics and Mathematics in
Biology, edited by T. A. Bancroft, J. W. Gowen and J. L. Lush.
Iowa State College Press, Ames, IA.
Dizon, A. E., B. L. Taylor and G. M. OCorry-Crowe, 1995 Why
statistical power is necessary to link analyses of molecular variation
to decisions about population structure. Am. Fish. Soc. Symp.
17: 288294.
Falconer, D. S., and T. F. C. Mackay, 1996 Introduction to Quantitative Genetics, Ed. 4. Longman, Essex, UK.
Hagen, R. L., 1997 In praise of the null hypothesis statistical test.
Am. Psychol. 52: 1524.
Harada, Y., M. Yokota and M. Iizuka, 1998 Genetic risk of domestication in artificial stocking and its possible reduction. Res. Popul.
Ecol. 40: 311324.
Hilborn, R., 1992 Hatcheries and the future of salmon in the Northwest. Fisheries 17: 58.
Hilborn, R., and J. Winton, 1993 Learning to enhance salmon
production: lessons from the salmonid enhancement program.
Can. J. Fish. Aquat. Sci. 50: 20432056.
Hill, W. G., 1979 A note on effective population size with overlapping generations. Genetics 92: 317322.
Hotta, T., T. Matsuishi, H. Sakano and Y. Kanno, 1999 Population structure of Pacific herring Clupea pallasii in the Eastern
Hokkaido waters. Nippon Suisan Gakkaishi. 64: 655660 (in Japanese).
Johnson, N. L., and S. Kotz, 1969 Discrete Distributions. Wiley, New
York.
Jorde, P. E., and N. Ryman, 1995 Temporal allele frequency change
and estimation of effective size in populations with overlapping
generations. Genetics 139: 10771090.
Kaeriyama, M., 1999 Hatchery programmes and stock management
of salmonid populations in Japan, pp. 153167 in Stock Enhancement and Sea Ranching, edited by B. R. Howell, E. Moksness
and T. Svasand. Blackwell, Oxford.
Kitada, S., 1999 Effectiveness of Japans stock enhancement programs: current perspectives, pp. 103131 in Stock Enhancement
and Sea Ranching, edited by B. R. Howell, E. Moksness and T.
Svasand. Blackwell, Oxford.
Kitada, S., K. Hiramatsu and H. Kishino, 1994 Estimating mortality rates from tag recoveries: incorporating over-dispersion, correlation and change points. ICES J. Mar. Sci. 51: 241251.
Knapp, G., 1999 Alaska salmon ranching: an economic review of
the Alaska salmon hatchery programme, pp. 537556 in Stock
Enhancement and Sea Ranching, edited by B. R. Howell, E. Moksness and T. Svasand. Blackwell, Oxford.
Krimbas, C. B., and S. Tsakas, 1971 The genetics of Dacus oleae.
V. Changes of esterase polymorphism in a natural population
following insecticide control: Selection of drift? Evolution 25:
454460.
Lange, K., 1995 Application of the Dirichlet distribution to forensic
match probabilities. Genetica 96: 107117.
Lee, P. M., 1989 Bayesian Statistics: An Introduction. Edward Arnold,
London.
Maritz, J. S., and T. L. Lwin, 1989 Empirical Bayes Methods, Ed. 2.
Chapman & Hall, London.
McCullagh, P., and J. A. Nelder, 1983 Generalized Linear Model,
Ed. 2. Chapman & Hall, London.
Miller, L. M., and A. R. Kapuscinski, 1997 Historical analysis of
genetic variation reveals low effective population size in a Northern Pike (Esox lucius) population. Genetics 147: 12491258.
Nei, M., 1987 Molecular Evolutionary Genetics. Columbia University
Press, New York.

Nei, M., and F. Tajima, 1981 Genetic drift and estimation of effective
population size. Genetics 98: 625640.
Peterman, R. M., 1990 Statistical power analysis can improve fisheries research and management. Can. J. Fish. Aquat. Sci. 47: 215.
Pollak, E., 1983 A new method for estimating the effective population size from allele frequency changes. Genetics 104: 531548.
Rannala, B., and J. A. Hartigan, 1996 Estimating gene flow in
island populations. Genet. Res. 67: 147158.
Ritter, J. A., 1997 The contribution of Atlantic salmon Salmo salar
L. enhancement to a sustainable resource. ICES J. Mar. Sci. 54:
11771187.
Roff, D. A., and P. Bentzen, 1989 The statistical analysis of mitochondrial DNA polymorphisms: 2 and the problem of small
sample. Mol. Biol. Evol. 6: 539545.
Sanghvi, L. D., 1953 Comparison of genetical and morphological
methods for a study of biological differences. Am. J. Phys. Anthropol. 11: 385404.
Snedecor, G. W., and W. G. Cochran, 1967 Statistical Methods, Ed.
6. Iowa State University Press, Ames, IA.
Tabata, K., and A. Mizuta, 1997 RFLP analysis of the mtDNA
D-loop region in red sea bream Pagrus major population from
four locations of western Japan. Fish. Sci. 63: 211217.
Tajima, F., 1992 Statistical method for estimating the effective population size in Pacific salmon. J. Hered. 83: 309311.
Utter, F., 1998 Genetic problems of hatchery-reared progeny released into the wild, and how to deal with them. Bull. Mar. Sci.
62(2): 623640.
Walters, C. J., 1988 Mixed-stock fisheries and the sustainability
of enhancement production for chinook and coho salmon, pp.
109115 in Salmon Production, Management, and Allocation, edited
by W. J. McNeil. Oregon State University Press, Corvallis, OR.
Waples, R. S., 1989 A general approach for estimating effective
population size from temporal changes in allele frequencies.
Genetics 121: 379391.
Waples, R. S., 1990 Conservation genetics of Pacific salmon: III.
Estimating effective population size. J. Hered. 81: 277289.
Waples, R. S., 1991 Genetics interactions between hatchery and wild
salmonids: lessons from the Pacific Northwest. Can. J. Fish. Aquat.
Sci. 48(Suppl. 1): 124133.
Waples, R. S., 1999 Dispelling some myths about hatcheries. Fisheries 24(2): 1221.
Weir, B. S., 1996 Genetic data analysis II. Sinauer Associates, Sunderland, MA.
Williamson, E. G., and M. Slatkin, 1999 Using maximum likelihood to estimate population size from temporal changes in allele
frequencies. Genetics 152: 755761.
Wright, S., 1931 Evolution in Mendelian populations. Genetics 16:
97159.
Wright, S., 1945 The differential equation of the distribution of
gene frequencies. Proc. Natl. Acad. Sci. USA 31: 383389.
Wright, S., 1951 The general structure of populations. Ann. Eugen.
15: 323354.
Wright, S., 1969 Evolution and Genetics of Populations: The Theory of
Gene Frequencies, Vol. 2. University of Chicago Press, Chicago.
Yoshizawa, K., 1997 Inbreeding of cultured ayu from allozyme allele
frequencies, Heisei 8 nendo seitaikei hozengata syubyo seisan gijyutu
kaihatsu kenkyu seika no gaiyo, pp. 3038. Fishery Agency of Japan,
Tokyo (in Japanese).
Communicating editor: Z-B. Zeng

APPENDIX

Expectation of X when X follows a 2 distribution:


Let X be a random variable that follows a 2 distribution
with degree of freedom of k. We derive here the expectation of X. Generally, when X is continuous and has a
probability density function f(x), the expectation of g(X)
is given by
E[g(X)] g(x)f(x)dx.

Empirical Bayes for Frequencies

For our case, the expectation of X is calculated as


E[X] Xf(x)dx

1
x(k 1)/2 1ex/2dx.
2k/2(k/2)

Let x/2 y, and we have


E[X]

1
(2y)(k1)/21ey 2dy.
2k/2(k/2)

2079

Using the gamma function, which is given by

y(k1)/21eydy

k1
,
2

finally we have
E(X) 2

((k1)/2)
.
(k/2)

You might also like